You are on page 1of 13

Reac Kinet Mech Cat (2010) 101:4961

DOI 10.1007/s11144-010-0195-x

Mathematical modeling of multi-bed adiabatic reactor


for the Methanol-to-Olefin process

Hao Hu Weiyong Ying Dingye Fang

Received: 6 January 2010 / Accepted: 26 April 2010 / Published online: 11 June 2010
Akademiai Kiado, Budapest, Hungary 2010

Abstract A mathematical model of a multi-bed adiabatic reactor for the Metha-


nol-to-Olefin (MTO) process is established based on a lumping kinetic equation
with SAPO-34 catalyst. Influences of different process conditions are investigated.
Temperature plays a more important role in governing simulation results than other
factors do. The decrease of methanol conversion resulting from catalyst deactivation
could be reflected by changing the model parameter.

Keywords Methanol-to-olefins  SAPO-34 catalyst  Multi-bed adiabatic reactor 


Process condition

List of symbols
A, B Modulus coefficient
a, b, c, d, e Modulus coefficient
Cor Activity correction factor
Cp Specific heat, J mol-1 K-1
D Inner diameter of reactor, m
Ea Activation energy, J mol-1
k Reaction rate constant, mol g-1 h-1 kPa-1
l Operational bed height, m
N Mole flow rate, mol h-1
N/A The mole ratio of nitrogen to methanol
m Mass flow rate, g h-1
n Mole flow rate, mol h-1
p Pressure, MPa

H. Hu  W. Ying  D. Fang (&)


Engineering Research Center of Large Scale Reactor Engineering and Technology,
Ministry Education, State Key Laboratory of Chemical Engineering,
East China University of Science and Technology, 200237 Shanghai, China
e-mail: dyfang@ecust.edu.cn

123
50 H. Hu et al.

r Reaction rate, mol g-1 h-1


T Temperature, K
w Mass of catalyst, g
WHSV Weight hourly space velocity, h-1
X Methanol conversion, %
y Mole ratio, %
a Stoichiometric coefficient
q Density of catalyst, kg m-3
DHhf Standard molar formation enthalpy, J mol-1
DHR Molar reaction enthalpy, J mol-1
Subscripts
0 Initial
A Methanol
b Bed
c Coolant
E Ethylene
i Component
in Inlet
out Outlet
P Propylene
RH Rest of hydrocarbons
T Total
W Water
Superscripts
Average

Introduction

The methanol-to-hydrocarbons reaction was first discovered in the laboratory of


Mobil Company using a ZSM-5 catalyst [1]. Higher temperature and lower pressure
were proved to have positive effects on the olefin generation, namely Methanol-to-
Olefins (MTO) process [2, 3]. SAPO-34, a micropore catalyst invented by Lok [4]
had shown substantial catalytic behavior during the MTO reaction. The UOP
Company promoted the MTO process in 1990s on this catalyst with the Norsk
Hydro Company [5, 6].
The mechanism of MTO, especially the route of the first CC bond generation,
has been long disputed by many scholars [5]. A parallel mechanism named
Hydrocarbon Pool Mechanism has been widely accepted, while the nature of the
Pool substance was assumed to be polymethylbenzene. Despite the fierce
arguments about the elementary studies, the commercialization of the MTO process
started much earlier. Kinetic studies were carried out by many scholars in order to
provide the fundamental data for scaling up the reactor. Bos [7] advanced his

123
Mathematical modeling of multi-bed adiabatic reactor 51

lumping kinetic model in terms of SAPO-34 catalyst in 1995. Coke was introduced
in his model, and was proved to have a positive effect for ethylene production [8].
Adopting Boss kinetic results, Soundararajan performed a circulating fluidized
reactor simulation, and investigated different process conditions. Gayubo [9]
introduced a more simplified reaction model, with which water addition for
deactivation attenuation was studied. Alwahabi [10] developed very detailed
reaction kinetics based on the single-event theory in terms of SAPO-34 catalyst with
the consideration of the coke formation which was assumed to be the result of large
molecules trapped inside the catalyst cavities. Simulation results on different types
of commercial scale reactors were compared as well [11] based on the detailed
kinetic model. Chen [12] reported his recent research about the MTO reaction and
deactivation kinetics using a special TEOM reactor.
SAPO-34 experiences a fast coking stage soon after the reaction initiates [13, 14].
The fast deactivation accompanying the high exothermicity of reaction has become
the chief consideration governing reactor selection [7]. Circulated fluidized bed was
proposed as a good choice for the MTO process [7, 8]. However, the relatively low
methanol conversion after a single pass with notable catalyst attrition, including the
high investment for scaling up the reactor, are the primary disadvantages [11].
A packed bed reactor is simple in construction and easily operated, which is an
elementary type of reactor to be considered. Unfortunately, research on the
influences of process conditions on the MTO reaction in such a reactor has not been
reported.

Reaction network and kinetic model

The products of the MTO reaction are quite complicated, including both
hydrocarbons and oxygenated compounds. A lumping kinetic model was hence
introduced in our previous study [15]. According to the Hydrocarbon Pool
Mechanism, all the products could be viewed as directly generated from oxygenated
feed (methanol and dimethyl ether) in a parallel way. All the DME products are
back calculated into methanol feed. The overall MTO reaction can be written as
Eq. 1 accordingly, while RH is the abbreviation of the rest of the hydrocarbons
(including C5? and alkanes), which is similar to Gayubo [9].
So far, reports of the establishment of the MTO plant with large capacity have not
been disclosed. However, performing a reactor simulation of the MTO process is of
interest in both theoretical and practical aspects. In this paper, referring to a lumping
kinetic model developed by our previous work [15], a multi-bed adiabatic reactor
with intermediate heat exchangers is selected for such a simulation, while the effects
of process conditions are investigated as well.
CH3 OH ! a1 CH4 a2 C2 H4 a3 C3 H6 a4 C4 H8 a5 RH H2 O 1
where ai is the stoichiometric coefficient.
The relationship between reaction enthalpy and temperature can be described by
Kirchhoff formula (Eq. 2).

123
52 H. Hu et al.

Z T
DHR T a1 DHhf;CH4 298 K CpCH4 dt a2 DHhf;C2 H4 298 K
298
Z T Z T
CpC2 H4 dt a3 DHhf;C3 H6 298 K CpC3 H6 dt
298 298
Z T
a4 DHhf;C4 H8 298 K CpC4 H8 dt a5 DHhf;RH 298 K
298
Z T Z T
h
CpRH dt DHf;H2 O 298 K Cpw dt  DHhf;CH4 O 298 K
298 298
Z T
CpA dt: 2
298

The adiabatic temperature rise is 450 K or so under a 648.2 K inlet temperature


from the calculation with Eq. 2. Therefore, multi-bed adiabatic reactor is necessary
for such a highly exothermic reaction so that the catalyst bed could be operated in an
allowable temperature.
Although the coking rate is high, it is noticed that the methanol conversion and
product distribution could stay relatively stable in a couple of hours after the
reaction initiates [7]. The catalyst turns into fast deactivation stage afterwards,
which gives a notable decrease of oxygenated conversion. Because a first order
global reaction rate is found [7, 9, 12], the kinetic model of methanol consumption
is proposed as Eq. 3, as the reaction remains in a stationary stage.
dNA EaA
rA  kA pA kA0 e RT p0 yA 3
dw
Here, k0A = 9.55 9 103 mol g cat-1 kPa-1 h-1, EaA = 5.24 9 104 J mol-1 [15],
the product generation rate follows the relationship below:
rA rCH4 rC2 H4 rC3 H6 rC4 H8 rRH rH2 O
: 4
1 a1 a2 a3 a4 a5 1
Parameter ai can be calculated through Eq. 5.
ai Ni =NT  X 5
It was found that ai changes with the reaction temperature, but had no obvious
relationship with space velocity in our previous study [15]. In fact, ai could be
viewed as the divided value between products yield and oxygenate conversion,
namely product selectivity. The parallel mechanism [5] also indicated the selectivity
is affected by the activation energy for each product, but has little relationship with
the feed concentration. An Arrhenius expression is found to give a best fit for each
parameter (Eq. 6).
ai Ai  expBi =T 6
All the values and the correlation coefficients in Eq. 6 are listed in Table 1 [15].

123
Mathematical modeling of multi-bed adiabatic reactor 53

Table 1 Values for different Ai and Bi


a1 a2 a3 a4 a5

Ai 297.61 11.97 0.028 1.40 9 10-4 4.96


Bi -7193.33 -3023.84 1142.09 4008.70 -4056.68
q2 0.9930 0.9979 0.9993 0.9967 0.9486

Reactor chosen and model formulation

Referring to the dimension of the packed bed reactors in the methanol synthesis
industry, a multi-bed adiabatic reactor with 6.0 m diameter is chosen in this paper.
Eq. 7 describes a mass balance relationship in the MTO process, which can be
deduced from Eq. 1.
P
1 yA0 ai
NT NT0 P 7
1 1 ai yA
The mathematical simulation for the multi-bed adiabatic reactor is developed from
the basic mass and heat balance equations of the pseudo-homogeneous one-
dimensional model, as shown in Eqs. 8 and 9. It is an elementary reactor model,
assuming that the gassolid phase reaction occurs in the homogenous phase. The
activity correction factor (Cor) is introduced into the model, which represents the
mass transfer effect, especially the internal diffusion coefficient of catalyst particles.
The introduction of Cor is an effective simplification of the reactor model, and the
radial dispersion inside each catalyst bed is viewed as negligible.
dNA pD2 qb
 Cor  rA 8
dl 4
dTb DHR dNA
 : 9
dl NT  Cp dl
All the fundamental data of catalyst properties and reactor geometry are listed in
Table 2.
The initial conditions for such ordinary differential equations are listed below:
When l = 0, NA NA0 ; NCH4 NC2 H4 NC3 H6 NC4 H8 NRH NH2 O 0;
Tb Tin : These ordinary differential equations are solved by the fourth-order
Runge-Kutta integration method. The temperature and concentration profiles in
different axial directions of the catalytic bed can thus be obtained. A flow rate of
2,700 kt/a pure methanol is selected as feedstock initially.

Table 2 Fundamental data for the catalyst and the reactor


Reactor Catalyst dimension Bed density Void
diameter (m) (mm) (kg m-3) fraction

6.0 4.2 9 4.2 640.0 0.5

123
54 H. Hu et al.

Temperature range of 648.2748.2 K is selected for each bed as typical operation


conditions. The restricted temperature for each bed is 773.2 K under every
circumstance because higher temperature fastens the coking rate as well as
byproduct generation.
Lower pressure favors the light olefin products. In this work, an operational
pressure of 0.3 MPa is selected. It is reported that the pressure drop inside the
reactor is negligible [11]. The equation of state for ideal gases could be adopted
under such conditions.

Results and discussions

Results at typical process condition

Calculation results indicate that a 4.5-m catalytic bed is required to reach a 300 kt/a
ethylene capacity under 2,700 kt/a methanol feed flow rate under the typical
reaction conditions. A volume of 127.2 m3 SAPO-34 catalyst is needed, which
amounts to 81.4 t. The weight hourly space velocity is 4.1 h-1 accordingly. Three
catalytic beds are required for the sake of removing reaction heat, with length of 1.0,
1.5 and 2.0 m respectively. The material balance after a single pass is listed as
Table 3, while profiles of temperature and concentrations are illustrated in Table 4.
The overall methanol conversion is 65.4%, with a mass ratio value of 1.0 for
ethylene and propylene. Take the 177 kt/a converted methanol as an example, 30 kt/a
ethylene together with 30 kt/a propylene are yielded, which is similar to the results
from the established MTO pilot plant (180 t converted methanol yields a 30 t
ethylene and 30 t propylene).

Effect of space velocity on simulation results

The space time is an important factor affecting operation results in practical use,
while it can be reflected by changing the space velocities (WHSV). In this study,
WHSV is ranged between 3.5 and 4.5 h-1, in order to avoid fast catalyst
deactivation. The inlet temperature and pressure for each bed are identical to the
reaction conditions mentioned above. Figs. 1 and 2 illustrate the trend of methanol
conversion and outlet temperature in each bed as WHSV changes.
Fig. 2 shows that the change of outlet temperature for each bed is complex as
WHSV changes. From the basic heat balance in Eq. 9, the increase of bed
temperature can be viewed as two sequential steps, namely heat release and heat
absorption. Under the condition of identical feed concentration (pure methanol), the

Table 3 Material balance relationship simulated from the multi-bed adiabatic reactor after a single pass
Feed (10 kt/a) Products (10 kt/a)

CH3OH CH4 C2H4 C3H6 C4H8 RH H2O Total

270.0 93.0 1.4 30.3 29.7 8.5 8.2 98.9 270.0

123
Mathematical modeling of multi-bed adiabatic reactor 55

Table 4 Temperature and concentration profiles in the multi-bed adiabatic reactor


Bed L Tb XA yCH4 yC2 H4 yC 3 H 6 yC4 H8 yRH yW
(m) (K) (%) (mol%) (mol%) (mol%) (mol%) (mol%) (mol%)

I 0 648.2 0 0 0 0 0 0 0
0.5 691.9 8.97 0.10 1.45 1.18 0.33 0.24 8.67
1.0 754.9 23.88 0.36 4.34 2.74 0.57 0.70 21.80
II 1.1 653.2 24.90 0.37 4.47 2.86 0.62 0.72 22.65
1.7 689.0 32.51 0.44 5.49 3.70 0.85 0.89 28.81
2.3 735.6 43.52 0.59 7.25 4.70 1.02 1.17 37.11
2.5 753.0 48.10 0.68 8.06 5.06 1.05 1.29 40.34
III 2.6 651.2 48.74 0.68 8.12 5.12 1.08 1.31 40.79
3.4 678.6 54.67 0.72 8.77 5.70 1.25 1.42 44.90
4.2 710.9 62.17 0.80 9.73 6.32 1.38 1.57 49.86
4.5 723.9 65.39 0.84 10.07 6.66 1.41 1.64 51.90

Fig. 1 Effect of WHSV on the 90


methanol conversion in each bed Bed I
80 Bed II
Bed III
70

60
XA(%)

50

40

30

20

10
3.6 3.8 4.0 4.2 4.4
-1
WHSV(h )

initial reaction rates are identical under every WHSV at the inlet of bed I. However,
lower space velocity means a smaller mole flow rate, hence favors a higher
temperature rise for such an adiabatic bed. The hot spot appeared at the outlet in
Bed I, which is commonly accepted for adiabatic bed. However, under a WHSV of
3.5 h-1, the temperature reaches 781.8 K at the outlet in Bed I, exceeds the limited
operation temperature.
The initial methanol concentration is not identical for different WHSV at the inlet
of Bed II, because of the different reaction results in Bed I. The methanol
concentration is lower for a smaller WHSV, and reduces the reaction rate and heat
release. However, the smaller mole flow rate of reaction mixture favors the heat
adsorption and the temperature rises. The competition of these two steps results in
the completely different trend at the outlet of Bed II for different space velocities.

123
56 H. Hu et al.

Fig. 2 Effect of WHSV on the Bed I


790
outlet temperature in each bed Bed II
780 Bed III

770

760

Tout(K)
750

740

730

720

3.6 3.8 4.0 4.2 4.4


WHSV(h-1)

Comparing with the first two beds, the decrease of initial methanol feed has a
greater effect on the bed temperature rise. Therefore, the temperature at the outlet of
Bed III rises as the WHSV increases, which is wholly opposite to the trend of Bed I.
The distribution of hydrocarbon products can be calculated as well (Fig. 3). The
concentrations for all the hydrocarbon compounds are almost constant under every
WHSV. According to the parallel reaction mechanism, all the products behave as
first-order reactions and generated from the oxygenated feed. Therefore, the
distribution of products relates to the distribution of reaction temperature more
closely. However, although with a change of outlet temperature for each catalyst
bed, all the temperature distribution at the axial profile under different WHSV
experiences from a lower value to higher value process. This could well explain the
simulation results.

Fig. 3 Effect of WHSV on the


product distribution at reactor 50
outlet

40
Mole Ratio(%)

30
Methane
Ethylene
20 Propylene
Butylene
Rest of Hydrocarbons
10

0
3.6 3.8 4.0 4.2 4.4
-1
WHSV(h )

123
Mathematical modeling of multi-bed adiabatic reactor 57

Effect of inlet temperature on simulation results

The inlet temperature for each bed is changed between the 623.2 and 773.2 K, as
relating to the reaction temperature of the MTO commercial plant from UOP/Hydro
(623.2773.2 K). Figs. 4 and 5 depict the effect of inlet temperature on methanol
conversion and outlet temperature in each bed.
The MTO reaction is sensitive to temperature, while higher reaction rate results
in a faster heat release and accelerates the reaction consequently. Fig. 4 illustrates
that the methanol conversion at the outlet of each bed increases following the rise of
inlet temperature. However, temperature runaway will occur if the inlet temperature
is too high.
The distribution of hydrocarbon products at the reactor outlet varies sharply with
the inlet temperature (Fig. 6). The selectivity for ethylene and by-products increases

Fig. 4 Effect of inlet Bed I


temperature on the methanol
75 Bed II
conversion in each bed
Bed III

60
XA(%)

45

30

15

620 630 640 650 660 670


Tin(K)

Fig. 5 Effect of inlet 820


temperature on the outlet Bed I
temperature in each bed 800 Bed II
Bed III
780

760
Tout(K)

740

720

700

680
620 630 640 650 660 670
Tin(K)

123
58 H. Hu et al.

Fig. 6 Effect of inlet 60


temperature on the product
distribution at the reactor outlet
50

40

Mole Ratio(%)
30 Methane
Ethylene
Propylene
20 Butylene
Rest of Hydrocarbons
10

0
620 630 640 650 660 670
Tin(K)

Table 5 Effect of inlet temperature on the value of ethylene/propylene


Tin (K) 623.2 633.2 643.2 653.2 663.2 673.2

mE/mp 0.78 0.88 0.97 1.08 1.18 1.30

Table 6 Comparison of simulation results in this study and previous reports


Reactor Inlet CH4 C2H4 C3H6 C4H8 RH
type temperature (mol%) (mol%) (mol%) (mol%) (mol%)
(K)

This study Packed bed 673.2 5.19 54.10 27.84 4.47 8.40
Alwahabi [11] Packed bed 648.2 3.97 49.03 32.20 8.70 6.10
UOP/Hydro [16] Fluidized bed 698.2 9.13 57.67 23.60 4.30 5.30
Soundararajan [8] Fluidized bed 723.2 11.76 49.96 25.12 7.07 6.08

as the temperature increases, while propylene and butylene selectivity decreases.


Accordingly, the mass ratio of ethylene to propylene varies from 0.78 to 1.30 as
temperature rises, as shown in Table 5, which is close to the results from circulated
fluidized bed reactor of UOP/Hydro (0.751.50). Based on varied reaction
temperatures and reactor types, different MTO results between this study and
previous reports are compared in Table 6. It can be seen that the simulation results
are close to the real data from commercial plant. Higher temperature yields a high
ethylene product, which proves that this simulation is reasonable.

Effect of feed composition on simulation results

Nitrogen and water are also selected as co-feed during the MTO process in practical
use. Because of the deactivation attenuation effect, co-feeding water proved to be a

123
Mathematical modeling of multi-bed adiabatic reactor 59

Fig. 7 Effect of feed 90


composition on the methanol Bed I
conversion in each bed 80 Bed II
Bed III
70

60

X (%)
50

A
40

30

20

10
0.0 0.1 0.2 0.3 0.4 0.5 0.6
N/A

Fig. 8 Effect of feed


800 Bed I
composition on the outlet
temperature in each bed 790
Bed II
Bed III
780

770
Tout (K)

760

750

740

730

0.0 0.1 0.2 0.3 0.4 0.5 0.6


N/A

good choice and attracted the attention of many scholars [9, 10]. Alwahabi [11] also
gave a detailed discussion focusing on this issue. In this work, nitrogen is chosen as
diluent co-fed with pure methanol.
Different amounts of nitrogen are added into methanol feed, while the 673.2 K inlet
temperature and 4.3 h-1 WHSV reaction conditions are selected as mentioned above.
Figs. 7 and 8 depict the variation of methanol conversion and outlet temperature in
each bed when changing the mole ratio of nitrogen to methanol feed (N/A).
It is evident that the addition of nitrogen alters both the MTO results and the
temperature distribution. However, it is also revealed that the value of mass ratio of
ethylene to propylene decreases slightly as the amount of nitrogen addition varies
from 0.05 to 0.55 (Table 7). For the sake of a high ethylene yield for the MTO
process, the addition of nitrogen can guarantee the normal operation while higher
inlet temperature is adopted.

123
60 H. Hu et al.

Table 7 Effect of feed composition on the mass ratio of ethylene to propylene


N/A 0.05 0.15 0.25 0.35 0.45 0.55

mE/mp 1.27 1.22 1.19 1.16 1.13 1.11

Effect of catalyst deactivation

The activity correction factor (Cor) represents the mass transfer process in
commercial scale plants, while the value is selected as 0.5 before catalyst
deactivation takes place. In fact, the decrease of catalyst activity due to coking
occurs in the whole MTO process. Coking generation will definitely increase the
mass transfer resistance inside the catalyst, as mentioned by Chen et al. [17].
However, obvious catalyst deactivation could only be detected after a certain
amount of coke is deposited already. Referring to the kinetic results in the early
study [7], the methanol conversion would reduce to nearly 50% in a quite short time
when the obvious deactivation stage starts. Therefore, the value of Cor is adjusted to
0.3 when deactivation is apparent in this study. The comparison between our results
and previous studies considering catalyst deactivation are listed in Table 8.
The two simulation results are contiguous when the reaction occurs at fresh
SAPO-34 catalyst. However, the change of Cor can represent the decrease of
methanol conversion caused by deactivation in multi-bed adiabatic reactor, while
the variation of hydrocarbon distribution cannot be reflected. This results from the
selective deactivation phenomenon for SAPO-34 catalyst [12].
It is concluded that the influence of selective deactivation should be taken into
consideration when serious deactivation occurs. Therefore, the relative data of
carbon deposition and catalyst deactivation are required. Besides, because of the fast
deactivation of SAPO-34 catalyst, regeneration should be performed simultaneously
while obvious deactivation is observed. Hence, at least one more plant with
identical construction is needed in parallel for the switch of reaction and
regeneration mode. The optimization of choosing the adequate time for the switch
of the reaction and regeneration mode is another complicate problem, which is
beyond the scope of this study.

Table 8 Comparison of simulation results for the MTO reaction considering deactivation
XA (%) Hydrocarbon distribution (mol%)

CH4 C2H4 C3H6 C4H8 RH

Before deactivation
This study 63.7 4.02 49.11 32.48 6.95 7.44
Alwahabi 62.5a 3.97 49.03 32.20 8.70 6.10
After deactivation
This study 42.3 3.50 46.55 33.95 8.29 7.68
Alwahabi 40.5a 15.45 48.25 26.67 5.88 3.74
a
DME in the simulation results is transformed to methanol feed

123
Mathematical modeling of multi-bed adiabatic reactor 61

Conclusions

A multi-bed adiabatic reactor equipped with intermediate heat exchangers for the
MTO process at the initial reaction stage is introduced in this study. A pseudo-
homogeneous one-dimensional mathematical model is established and solved by the
fourth-order Runge-Kutta integration method. Under the condition of a mass flow
rate of 2,700 kt/a methanol feed, a 4.5 m catalytic bed with 6.0 m diameter is
required for a 300 kt/a ethylene capacity. The effects of WHSV, inlet temperature,
feed composition and catalyst deactivation are discussed.
The inlet temperature in each bed is the most important factor governing the
MTO process. The addition of nitrogen as diluent has a positive impact on the bed
temperature control especially at high inlet temperatures. Catalyst deactivation
changes methanol conversion as well as product selectivity. Adjusting model
parameter Cor can represent the variation of methanol conversion caused by the loss
of activity by comparison with previous results. However, data on deactivation
kinetics are required when considering the variation of products distribution caused
by deactivation as well, and could be solved by further investigations. As the
catalyst can be regenerated by burning off coke in the original plant, at least one
more plant is needed in parallel for the switch of reaction and regeneration mode.
In general, by comparison with simulation results of previous reports, this paper
gives a reasonable description of the behavior of the MTO process in a multi-bed
adiabatic reactor based on the simple and practical lumping kinetic model, while the
catalyst activity is at the initial state.

References

1. Meisel SL, McCullough JP, Lechthaler CH (1976) Chemtech 6:86


2. Chang CD, Lang WH, Smith RL (1979) J Catal 56:169
3. Chang CD, Chu TW, Socha RF (1984) J Catal 86:289
4. Lok BM, Messina CA (1984) J Am Chem Soc 106:6092
5. Stocker M (1999) Microporous Mesoporous Mater 29:3
6. Keil FJ (1999) Microporous Mesoporous Mater 29:49
7. Bos AN, Tromp PJ, Akse HN (1995) Ind Eng Chem Res 34:3808
8. Soundararajan S, Dalai AK, Berruti F (2001) Fuel 80:1187
9. Gayubo AG, Aguayo AT, Sanchez del Campo AE, Tarrio AM, Bilbao J (2000) Ind Eng Chem Res
39:292
10. Alwahabi SM, Froment GF (2004) Ind Eng Chem Res 43:5098
11. Alwahabi SM, Froment GF (2004) Ind Eng Chem Res 43:5112
12. Chen D, Grnvold A, Moljord K, Holmen A (2007) Ind Eng Chem Res 46:4116
13. Chen D, Rebo HP, Grnvold A, Moljord K, Holmen A (2000) Microporous Mesoporous Mater
3536:121
14. Qi GZ, Xie ZK, Yang WM, Zhong SQ, Liu HX, Zhang CF, Chen QL (2007) Fuel Process Technol
88:437
15. Hu H, Ye L, Ying W, Fang D (2009) J East China Univ Sci Technol (Nat Sci Ed) 5:655
16. Gao JS, Zhang DX (2006) Chin Coal Chem Ind 4:7
17. Chen D, Rebo HP, Moljord K, Holmen A (1997) Ind Eng Chem Res 36:3473

123

You might also like